This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Half-plane diffraction problems on a triangular lattice

D. Kapanadze [email protected] E. Pesetskaya [email protected] A. Razmadze Mathematical Institute, TSU, Merab Aleksidze II Lane 2, Tbilisi 0193, Georgia Free University of Tbilisi, Tbilisi 0159, Georgia
Abstract

We investigate thin-slit diffraction problems for two-dimensional lattice waves. The peculiar structure allows us to consider the problems on the semi-infinite triangular lattice, consequently, we study Dirichlet problems for the two-dimen-sional discrete Helmholtz equation in a half-plane. In view of the existence and uniqueness of the solution, we provide new results for the real wave number k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\} without passing to the complex wave number and derive an exact representation formula for the solution. For this purpose, we use the notion of the radiating solution. Finally, we propose a method for numerical calculation. The efficiency of our approach is demonstrated in an example related to the propagation of wave fronts in metamaterials through two small openings.

keywords:
discrete Helmholtz equation, Dirichlet boundary value problem, half-plane diffraction, metamaterials, triangular lattice model
journal: arXiv.org

1 Introduction

Wave propagation through discrete structures remains an active area of research today. The triangular lattice is one of the five two-dimensional Bravais lattice types and appears naturally in applications. For example, close-packed planes occur frequently in some kinds of crystals in the form of triangular lattice [2, 3]. Besides, one can consider two-dimensional passive propagation media, a host microstrip line network periodically loaded with series capacitors and shunt inductors for signal processing and filtering (as shown in Figure 1). This type of inductor-capacitor lattice is referred to a negative-refractive-index transmission-line (NRI-TL) metamaterial [6] or, simply, left-handed 2D metamaterial.

Motivated by applications of recent interest related to analog circuits, crystalline materials, and metamaterials, we investigate thin-slit diffraction problems for triangular lattices. The peculiar structure allows us to consider the problems on the semi-infinite triangular lattice. For this lattice, we study Dirichlet problems for the two-dimensional discrete Helmholtz equation in a half-plane (mathematically formulated in the next section). When one is interested in the analysis of regular processes in which waves corresponding to the microstructural scales can be neglected, then the continuum limit of corresponding equations can be investigated. In this case, we arrive at the famous problem in applied mathematics – the half-plane diffraction problem. It is well known that the classical continuum model of wave diffraction can be considered only as the slowly-varying approximation of a discrete or structured material. Nowadays the understanding of nanostructure and microstructure phenomena in modern materials and composites in critical conditions is of crucial importance. It turns out that a discrete structure provides an effective way to describe microstructural processes and influences within, cf., e.g., [5, 6, 13, 17]. Therefore, there is an obvious necessity to avoid continuum limits and instead directly analyze the discrete Helmholtz diffraction problems. Although the similar problem for square lattice has been studied in [1], see also [15], its extension to a triangular lattice model is not direct.

Our main interest lies in the investigation of the problems with wave number kk within the pass-band which is an arbitrary non-zero complex number in general. In this paper, we use the well approved approach to provide new results for a real wave number k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\} without passing to a complex wave number. In order to have the unique solvability results and an exact representation formula for the solution, we use the notion of the radiating solution [14] and the method of images to construct the Dirichlet Green’s function for the half-plane. Notice that the cases with the wave numbers within the stop-band is mathematically more simple, they do not need the radiation conditions and is not considered here. Finally, we propose a method for numerical calculation. The efficiency of our approach is demonstrated in an example related to the propagation of wave fronts in metamaterials through two small openings.

2 Basic notations and formulation of the problem

Following the customary notation in mathematics, let \mathbb{Z}, +\mathbb{Z}^{+}, \mathbb{Z}^{-}, \mathbb{N}, \mathbb{R}, and \mathbb{C} denote the set of integers, positive integers, negative integers, non-negative integers, real numbers and complex numbers, respectively. We denote by e1=(1,0)e_{1}=(1,0), e2=(0,1)e_{2}=(0,1) the standard base of 2\mathbb{Z}^{2} (=×=\mathbb{Z}\times\mathbb{Z}).

Consider a two-dimensional infinite triangular lattice 𝔗\mathfrak{T} defined as a periodic simple graph {𝒱,}\{\mathcal{V},\mathcal{E}\}, where

𝒱={(x1+x2/2,3x2/2)2:(x1,x2)2}\mathcal{V}=\{(x_{1}+x_{2}/2,\sqrt{3}x_{2}/2)\subset\mathbb{R}^{2}:(x_{1},x_{2})\in\mathbb{Z}^{2}\}

is a vertex set, and \mathcal{E} is an edge set, whose endpoints (a,b)𝒱×𝒱(a,b)\in\mathcal{V}\times\mathcal{V} are adjacent points, i.e., ab=1\mid a-b\mid=1, cf. Figure 1. The time-harmonic discrete waves in 𝔗\mathfrak{T} can be described by solutions of the discrete Helmholtz equation

Refer to caption

Refer to caption

Figure 1: Triangular lattice. Connection between x=(x1,x2)2x=(x_{1},x_{2})\in\mathbb{Z}^{2} and the Euclidean coordinates of the vertexes (black dots) is established via (x1,x2)(x1+x2/2,3x2/2)(x_{1},x_{2})\to(x_{1}+x_{2}/2,\sqrt{3}x_{2}/2). The nearest neighbor interactions based on the triangular lattice are shown with thick blue lines. Triangular lattice can be viewed as a left-handed 2D inductor-capacitor metamaterial. A host transmission-line is loaded periodically with series capacitors and shunt inductors.
(Δd+k2)u(x)=g(x),x=(x1,x2)2.(\Delta_{d}+k^{2})u(x)=g(x),\quad x=(x_{1},x_{2})\in\mathbb{Z}^{2}. (1)

Here, Δd\Delta_{d} denotes the discrete (a 7-point) Laplacian defined as follows

Δdu(x)=\displaystyle\Delta_{d}u(x)= u(x+e1)+u(xe1)+u(x+e2)+u(xe2)\displaystyle u(x+e_{1})+u(x-e_{1})+u(x+e_{2})+u(x-e_{2})
+u(x+e1e2)+u(xe1+e2)6u(x),\displaystyle+u(x+e_{1}-e_{2})+u(x-e_{1}+e_{2})-6u(x),

where gg is a function with bounded support, i.e., gC0(2)g\in C_{0}(\mathbb{Z}^{2}).

Let Ω=Ω̊Γ\Omega=\mathring{\Omega}\cup\Gamma be a discrete half-plane in 2\mathbb{Z}^{2}, where Ω̊:=×+\mathring{\Omega}:=\mathbb{Z}\times\mathbb{\mathbb{Z}^{+}} and Γ:=Ω={(x1,0):x1}\Gamma:=\partial\Omega=\{(x_{1},0):x_{1}\in\mathbb{Z}\}. We consider the Dirichlet poblem for the discrete Helmholtz equation on the semi-infinite lattice:

(Δd+k2)u(x)\displaystyle(\Delta_{d}+k^{2})u(x) =0,inΩ̊,\displaystyle=0,\quad\quad\ \ \textup{in}\ \mathring{\Omega}, (2a)
u(y)\displaystyle u(y) =f(y),onΓ.\displaystyle=f(y),\quad\textup{on}\ \Gamma. (2b)

Here, ff is a given function supported on a finite subset of Γ\Gamma.

Thus, we are interested in studying the problem of the existence and uniqueness of a function u:Ωu:\Omega\to\mathbb{C} such that uu satisfies the discrete Helmoltz equation (2a) with k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\} and the boundary condition (2b). From now on we will refer to this problem as Problem 𝒫D\mathcal{P}_{\mathrm{D}}.

3 Lattice Green’s function

Denote by 𝒢(x;y)\mathcal{G}(x;y) the Green’s function for the discrete Helmholtz equation (1) centered at yy and evaluated at xx. Then the function 𝒢(x;y)\mathcal{G}(x;y) satisfies

(Δd+k2)𝒢(x;y)=δx,y,(\Delta_{d}+k^{2})\mathcal{G}(x;y)=\delta_{x,y}, (3)

where δx,y\delta_{x,y} is the Kronecker delta. For brevity, we use the notation 𝒢(x)\mathcal{G}(x) for 𝒢(x;0)\mathcal{G}(x;0). Notice that 𝒢(x;y)=𝒢(xy)\mathcal{G}(x;y)=\mathcal{G}(x-y).

Using the discrete Fourier transform and the inverse Fourier transform we get

𝒢(x)=14π2ππππeι(xξ)σ(ξ;k)𝑑ξ,ξ=(ξ1,ξ2),\mathcal{G}(x)=\frac{1}{4\pi^{2}}\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}\frac{e^{\iota(x\cdot\xi)}}{\sigma(\xi;k)}d\xi,\quad\xi=(\xi_{1},\xi_{2}), (4)

where

σ(ξ;k2)\displaystyle\sigma(\xi;k^{2}) =eιξ1+eιξ1+eιξ2+eιξ2+eιξ1eιξ2+eιξ1eιξ26+k2\displaystyle=e^{\iota\xi_{1}}+e^{-\iota\xi_{1}}+e^{\iota\xi_{2}}+e^{-\iota\xi_{2}}+e^{\iota\xi_{1}}e^{-\iota\xi_{2}}+e^{-\iota\xi_{1}}e^{\iota\xi_{2}}-6+k^{2} (5)
=k26+2cosξ1+2cosξ2+2cos(ξ1ξ2).\displaystyle=k^{2}-6+2\cos\xi_{1}+2\cos\xi_{2}+2\cos(\xi_{1}-\xi_{2}).

The lattice Green’s function 𝒢\mathcal{G} is quite well known when k2\[0,9]k^{2}\in\mathbb{C}\backslash[0,9] (cf., e.g., [16]). Notice that if k2\[0,9]k^{2}\in\mathbb{C}\backslash[0,9] then σ0\sigma\neq 0 and, consequently, 𝒢\mathcal{G} in (4) is well defined. In this case 𝒢(x)\mathcal{G}(x) decays exponentially when x\mid x\mid\to\infty. Moreover, we have

𝒢(x1,x2)=𝒢(x2,x1)=𝒢(x1,x2)=𝒢(x1+x2,x2)\mathcal{G}(x_{1},x_{2})=\mathcal{G}(x_{2},x_{1})=\mathcal{G}(-x_{1},-x_{2})=\mathcal{G}(x_{1}+x_{2},-x_{2}) (6)

for all x=(x1,x2)2x=(x_{1},x_{2})\in\mathbb{Z}^{2}.

Let us show 𝒢(x1,x2)=𝒢(x1+x2,x2)\mathcal{G}(x_{1},x_{2})=\mathcal{G}(x_{1}+x_{2},-x_{2}) while other identities are trivial. According to (4), we have

𝒢(x1+x2,x2)\displaystyle\mathcal{G}(x_{1}+x_{2},-x_{2}) =\displaystyle= 14π2ππππeιx1ξ1+ιx2(ξ1ξ2)σ(ξ1,ξ2;k)𝑑ξ1𝑑ξ2\displaystyle\frac{1}{4\pi^{2}}\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}\frac{e^{\iota x_{1}\xi_{1}+\iota x_{2}(\xi_{1}-\xi_{2})}}{\sigma(\xi_{1},\xi_{2};k)}d\xi_{1}d\xi_{2}
=\displaystyle= 14π2ππη1πη1+πeιx1η1+ιx2η2σ(η1,η2;k)𝑑η2𝑑η1\displaystyle\frac{1}{4\pi^{2}}\int_{-\pi}^{\pi}\int_{\eta_{1}-\pi}^{\eta_{1}+\pi}\frac{e^{\iota x_{1}\eta_{1}+\iota x_{2}\eta_{2}}}{\sigma(\eta_{1},\eta_{2};k)}d\eta_{2}d\eta_{1}

with η1=ξ1,η2=ξ1ξ2\eta_{1}=\xi_{1},\,\eta_{2}=\xi_{1}-\xi_{2}. The following equality can be easily obtained by changing the variable η2\eta_{2} to η22π\eta_{2}-2\pi

0ππη1+πeιx1η1+ιx2η2σ(η1,η2;k)𝑑η2𝑑η1=0ππη1πeιx1η1+ιx2η2σ(η1,η2;k)𝑑η2𝑑η1.\int_{0}^{\pi}\int_{\pi}^{\eta_{1}+\pi}\frac{e^{\iota x_{1}\eta_{1}+\iota x_{2}\eta_{2}}}{\sigma(\eta_{1},\eta_{2};k)}d\eta_{2}d\eta_{1}=\int_{0}^{\pi}\int_{-\pi}^{\eta_{1}-\pi}\frac{e^{\iota x_{1}\eta_{1}+\iota x_{2}\eta_{2}}}{\sigma(\eta_{1},\eta_{2};k)}d\eta_{2}d\eta_{1}.

The factor eιx22πe^{\iota x_{2}2\pi} is equal to 11 since x2x_{2}\in\mathbb{Z}. Similar arguments give us

π0η1ππeιx1η1+ιx2η2σ(η1,η2;k)𝑑η2𝑑η1=π0η1+ππeιx1η1+ιx2η2σ(η1,η2;k)𝑑η2𝑑η1.\int_{-\pi}^{0}\int_{\eta_{1}-\pi}^{-\pi}\frac{e^{\iota x_{1}\eta_{1}+\iota x_{2}\eta_{2}}}{\sigma(\eta_{1},\eta_{2};k)}d\eta_{2}d\eta_{1}=\int_{-\pi}^{0}\int_{\eta_{1}+\pi}^{\pi}\frac{e^{\iota x_{1}\eta_{1}+\iota x_{2}\eta_{2}}}{\sigma(\eta_{1},\eta_{2};k)}d\eta_{2}d\eta_{1}.

Taking into account the last two equalities, we get

𝒢(x1+x2,x2)=14π2ππππeιx1η1+ιx2η2σ(η1,η2;k)𝑑η2𝑑η1=𝒢(x1,x2).\mathcal{G}(x_{1}+x_{2},-x_{2})=\frac{1}{4\pi^{2}}\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}\frac{e^{\iota x_{1}\eta_{1}+\iota x_{2}\eta_{2}}}{\sigma(\eta_{1},\eta_{2};k)}d\eta_{2}d\eta_{1}=\mathcal{G}(x_{1},x_{2}).

In the case k2(0,9)\{8}k^{2}\in(0,9)\backslash\{8\} the expression (4) is understood as follows: we replace k2k^{2} by k2+ιεk^{2}+\iota\varepsilon with 0<ε10<\varepsilon\ll 1 and let ε0\varepsilon\to 0 at the end of the calculation, cf. [14]. Thus, we define the lattice Green’s function for k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\} as a pointwise limit of

(Rλ+ιεδx,0)(x):=14π2ππππeιxξdξ1dξ2σ(ξ;k2+ιε)(R_{\lambda+\iota\varepsilon}\delta_{x,0})(x):=\frac{1}{4\pi^{2}}\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}\frac{e^{\iota x\cdot\xi}d\xi_{1}d\xi_{2}}{\sigma(\xi;k^{2}+\iota\varepsilon)}

as k2+ιεk2+ι0k^{2}+\iota\varepsilon\to k^{2}+\iota 0 and denote it again by 𝒢(x)\mathcal{G}(x), i.e., 𝒢(x)=(Rλ+ι0δx,0)(x)\mathcal{G}(x)=(R_{\lambda+\iota 0}\delta_{x,0})(x). Clearly, 𝒢(x)\mathcal{G}(x) is a solution to equation (3) and satisfies equalities (6).

4 Unique solvability result

In order to simplify further arguments let us introduce following vectors:

e1=(1,0),e2=(0,1),e3=e1e2,e4=e1,e5=e2,e6=e3.\displaystyle e_{1}=(1,0),\quad e_{2}=(0,1),\quad e_{3}=e_{1}-e_{2},\quad e_{4}=-e_{1},\quad e_{5}=-e_{2},\quad e_{6}=-e_{3}.

For any point x2x\in\mathbb{Z}^{2} we define the 6-neighbourhood Fx0F^{0}_{x} as the set of points {x+ej:j=1,,6}\{x+e_{j}:j=1,...,6\} and the neighbourhood FxF_{x} as Fx0{x}F^{0}_{x}\bigcup\{x\}. We say that R2R\subset\mathbb{Z}^{2} is a region if there exist disjoint nonempty subsets R̊\mathring{R} and R\partial R of RR such that

  • (a)

    R=R̊RR=\mathring{R}\cup\partial R,

  • (b)

    if xR̊x\in\mathring{R} then FxRF_{x}\subset R,

  • (c)

    if xRx\in\partial R then there is at least one point yFx0y\in F^{0}_{x} such that yR̊y\in\mathring{R}.

Clearly, the subsets R̊\mathring{R} and R\partial R are not defined uniquely by RR, but henceforth, it will be always assumed that R̊\mathring{R} and R\partial R are also given and fixed for a given region RR in 2\mathbb{Z}^{2}. We also say that xx is an interior (boundary) point of RR if xR̊x\in\mathring{R} (xRx\in\partial R).

Denote by (R)j(\partial R)_{j}, j=1,,6j=1,...,6, a set of all boundary points yRy\in\partial R such that yejR̊y-e_{j}\in\mathring{R} and call it the sides of the boundary RR. Note that a boundary point xx can simultaneously belong to all six sides of RR. Thus, these sides may overlap each other. Clearly, R\partial R is the union of its six sides, R=j=16(R)j\partial R=\cup_{j=1}^{6}(\partial R)_{j}.

As it was already mentioned above, yRy\in\partial R may be a point of intersection of several sides of R\partial R. However, in our arguments presented below it will be always clear which side is needed to be considered. Under this condition, we define the discrete derivative in the outward normal direction eje_{j}, j=1,,6j=1,\dots,6,

Tu(y)=u(y)u(yej),y(R)j.Tu(y)=u(y)-u(y-e_{j}),\quad y\in(\partial R)_{j}.

Let us introduce the following set H0={(0,0)}H_{0}=\{(0,0)\} and then define HNH_{N}, NN\in\mathbb{N}, with the help of recurrence formula

HN:=xHN1FxH_{N}:=\bigcup_{x\in H_{N-1}}F_{x}

with H̊N:=HN1\mathring{H}_{N}:=H_{N-1} and (H)N:=HN\H̊N(\partial H)_{N}:=H_{N}\backslash\mathring{H}_{N}.

Let RR be a finite region. Representing R=xR̊FxR=\bigcup_{x\in\mathring{R}}{F_{x}}, we can easily derive a discrete analogue of Green’s second identity

xR̊(u(x)Δdv(x)v(x)Δdu(x))=yR(u(y)Tv(y)v(y)Tu(y)).\sum_{x\in\mathring{R}}(u(x)\Delta_{d}v(x)-v(x)\Delta_{d}u(x))=\sum_{y\in\partial R}(u(y)Tv(y)-v(y)Tu(y)). (7)

Now let us give a definition of a radiating solution on the discrete half-plane. First, we consider the case k2(0,8)k^{2}\in(0,8). We say that u:Ωu:\Omega\to\mathbb{C} satisfies the radiation condition at infinity if

{u(x)=O(x12),u(x+ej)=eιξj(α,k)u(x)+O(x32),j=1,2,\left\{\begin{aligned} u(x)&=O(\mid x\mid^{-\frac{1}{2}}),\\ u(x+e_{j})&=e^{\iota\xi^{*}_{j}(\alpha,k)}u(x)+O(\mid x\mid^{-\frac{3}{2}}),\quad j=1,2,\end{aligned}\right. (8)

with the remaining term decaying uniformly in all directions x/xx/\mid x\mid, where xx is characterized as x1=xcosαx_{1}=\mid x\mid\cos\alpha, x2=xsinαx_{2}=\mid x\mid\sin\alpha, 0απ0\leq\alpha\leq\pi. Here, ξj(α,k)\xi^{*}_{j}(\alpha,k) is the jjth coordinate of the point ξ(α,k)\xi^{*}(\alpha,k).

For k2(8,9)k^{2}\in(8,9) we say that u:Ωu:\Omega\to\mathbb{C} satisfies the radiation condition at infinity if it can be represented as a sum of functions u1u_{1} and u2u_{2} such that each of them satisfies

{ui(x)=O(x12),ui(x+ej)=eιξj,i(α,k)ui(x)+O(x32),i,j=1,2,\left\{\begin{aligned} u_{i}(x)&=O(\mid x\mid^{-\frac{1}{2}}),\\ u_{i}(x+e_{j})&=e^{\iota\xi^{*,i}_{j}(\alpha,k)}u_{i}(x)+O(\mid x\mid^{-\frac{3}{2}}),\quad i,j=1,2,\end{aligned}\right. (9)

with the remaining term decaying uniformly in all directions x/xx/\mid x\mid, where xx is characterized as x1=xcosαx_{1}=\mid x\mid\cos\alpha, x2=xsinαx_{2}=\mid x\mid\sin\alpha, 0απ0\leq\alpha\leq\pi. Here, ξj,i(α,k)\xi^{*,i}_{j}(\alpha,k) is the jjth coordinate of the point ξ,i(α,k)\xi^{*,i}(\alpha,k), i=1,2i=1,2.

It is shown in [14] that 𝒢(x)\mathcal{G}(x) satisfies the radiation conditions introduced above. Now we are ready to introduce a notion of a radiating solution to the discrete Helmholtz equation (1).

Definition 4.1

Let k2(0,8)k^{2}\in(0,8). A solution uu to the discrete Helmholtz equation (2a) is called radiating if it satisfies the radiation condition (8).

Let k2(8,9)k^{2}\in(8,9). A solution uu to the discrete Helmholtz equation (2a) is called radiating if it satisfies the radiation condition (9).

We represent the Dirichlet Green’s function for the half-plane as follows:

𝒢+(x;y)=𝒢(x;y)𝒢(x^;y),y2,\mathcal{G}^{+}(x;y)=\mathcal{G}(x;y)-\mathcal{G}(\hat{x};y),\quad y\in\mathbb{Z}^{2},

with x=(x1,x2)Ωx=(x_{1},x_{2})\in\Omega and x^=(x1+x2,x2)\hat{x}=(x_{1}+x_{2},-x_{2}). Notice that 𝒢+(x;y)\mathcal{G}^{+}(x;y) can be represented equivalently as

𝒢+(x;y)=𝒢(x;y)𝒢(x;y^).\mathcal{G}^{+}(x;y)=\mathcal{G}(x;y)-\mathcal{G}(x;\hat{y}).

Indeed, using the property (6), we have

𝒢(x^;y)=𝒢(x1+x2y1,x2y2)=𝒢(x1y1y2,x2+y2)=𝒢(x;y^).\mathcal{G}(\hat{x};y)=\mathcal{G}(x_{1}+x_{2}-y_{1},-x_{2}-y_{2})=\mathcal{G}(x_{1}-y_{1}-y_{2},x_{2}+y_{2})=\mathcal{G}(x;\hat{y}). (10)

Since (y1,0)^=(y1,0)\widehat{(y_{1},0)}=(y_{1},0), we immediately get 𝒢+(x;y)=0\mathcal{G}^{+}(x;y)=0 for yΓy\in\Gamma. Moreover,

(Δd+k2)𝒢+(x;y)=(Δd+k2)𝒢(x;y)(Δd+k2)𝒢(x^;y)=δx,yδx^,y=δx,y(\Delta_{d}+k^{2})\mathcal{G}^{+}(x;y)=(\Delta_{d}+k^{2})\mathcal{G}(x;y)-(\Delta_{d}+k^{2})\mathcal{G}(\hat{x};y)=\delta_{x,y}-\delta_{\hat{x},y}=\delta_{x,y}

for all xΩ̊x\in\mathring{\Omega} and yΩy\in\Omega. From (10) we see that 𝒢+(x;y)\mathcal{G}^{+}(x;y) satisfies the radiation condition. Indeed, for any fixed yΓy\in\Gamma the angle α\alpha from the radiation conditions (8) and (9) is the same for 𝒢(x;y)=𝒢(xy)\mathcal{G}(x;y)=\mathcal{G}(x-y) and 𝒢(x;y^)=𝒢(xy^)\mathcal{G}(x;\hat{y})=\mathcal{G}(x-\hat{y}) when x\mid x\mid\to\infty.

Theorem 4.2

Let k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\}, and uu be a given function Ω\Omega\to\mathbb{C} that satisfies the radiation condition. The function uΓu\mid_{\Gamma} has a finite support. Then, for any point xΩ̊x\in\mathring{\Omega}, we have the following representation formula

u(x)=yΓ(u(y)T𝒢+(x;y)𝒢+(x;y)Tu(y))+yΩ̊𝒢+(x;y)(Δd+k2)u(y).u(x)=\sum_{y\in\Gamma}\big{(}u(y)T\mathcal{G}^{+}(x;y)-\mathcal{G}^{+}(x;y)Tu(y)\big{)}+\sum_{y\in\mathring{\Omega}}\mathcal{G}^{+}(x;y)(\Delta_{d}+k^{2})u(y).

In particular, if uu is a solution to the discrete Helmholtz equation

(Δd+k2)u(x)=0inΩ̊,(\Delta_{d}+k^{2})u(x)=0\quad\textup{in}\ \mathring{\Omega},

then

u(x)=yΓu(y)T𝒢+(x;y).u(x)=\sum_{y\in\Gamma}u(y)T\mathcal{G}^{+}(x;y). (11)
  • Proof.  Denote by u~\tilde{u} the extension of uu to 2\mathbb{Z}^{2} by zeros, i.e, u~(x)=u(x)\tilde{u}(x)=u(x) if xΩx\in\Omega and u~(x)=0\tilde{u}(x)=0 if x×x\in\mathbb{Z}\times\mathbb{Z}^{-}. Then, for any finite region HNH_{N}, NN\in\mathbb{N}, we apply Green’s second identity (7) where we take v(y)=𝒢+(x;y)v(y)=\mathcal{G}^{+}(x;y). Here, note that 𝒢+\mathcal{G}^{+} is a fundamental solution for yΩ̊y\in\mathring{\Omega}, and u~(y)=0\tilde{u}(y)=0 for y×y\in\mathbb{Z}\times\mathbb{Z}^{-}. Thus, u(y)(Δd+k2)v(y)u(y)(\Delta_{d}+k^{2})v(y) disappears in the following identity

    u(y)Δdv(y)v(y)Δdu(y)=u(y)(Δd+k2)v(y)v(y)(Δd+k2)u(y)u(y)\Delta_{d}v(y)-v(y)\Delta_{d}u(y)=u(y)(\Delta_{d}+k^{2})v(y)-v(y)(\Delta_{d}+k^{2})u(y)

    as far as y×y\in\mathbb{Z}\times\mathbb{Z}^{-}. For xH̊Nx\in\mathring{H}_{N}, we get

    u~(x)\displaystyle\tilde{u}(x) =yH̊Nu~(y)δx,y\displaystyle=\sum_{y\in\mathring{H}_{N}}\tilde{u}(y)\delta_{x,y}
    =yHN(u~(y)T𝒢+(x;y)𝒢+(x;y)Tu~(y))+yH̊N(𝒢+(x;y)(Δd+k2)u~(y)).\displaystyle=\sum_{y\in\partial H_{N}}\big{(}\tilde{u}(y)T\mathcal{G}^{+}(x;y)-\mathcal{G}^{+}(x;y)T\tilde{u}(y)\big{)}+\sum_{y\in\mathring{H}_{N}}(\mathcal{G}^{+}(x;y)(\Delta_{d}+k^{2})\tilde{u}(y)).

    Passing to the limit NN\to\infty, we use exactly the same arguments as in the proof of Theorem 5.2 from [14] and, consequently, we get

    u~(x)=y2𝒢+(x;y)(Δd+k2)u~(y).\tilde{u}(x)=\sum_{y\in\mathbb{Z}^{2}}\mathcal{G}^{+}(x;y)(\Delta_{d}+k^{2})\tilde{u}(y).

    For the function uu, we obtain

    u(x)=yΓ(u(y)T𝒢+(x;y)𝒢+(x;y)Tu(y))+yΩ̊𝒢+(x;y)(Δd+k2)u(y).u(x)=\sum_{y\in\Gamma}\big{(}u(y)T\mathcal{G}^{+}(x;y)-\mathcal{G}^{+}(x;y)Tu(y)\big{)}+\sum_{y\in\mathring{\Omega}}\mathcal{G}^{+}(x;y)(\Delta_{d}+k^{2})u(y).

    Taking into the account that 𝒢+(x;y)=0\mathcal{G}^{+}(x;y)=0 when yΓy\in\Gamma, for a solution uu to the discrete Helmholtz equation we get the following quality

    u(x)=yΓu(y)T𝒢+(x;y),xΩ̊.u(x)=\sum_{y\in\Gamma}u(y)T\mathcal{G}^{+}(x;y),\quad x\in\mathring{\Omega}.

    \Box

Now we are ready to prove the unique solvability result for the discrete Helmholtz equation on the semi-infinite triangular lattice.

Theorem 4.3

Let k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\} then the Problem 𝒫D\mathcal{P}_{\mathrm{D}} has a unique radiating solution uu which can be represented as

u(x)=yΓ(δx,y𝒢+(x;y+e2)𝒢+(x,y+e2e1))f(y).u(x)=\sum_{y\in\Gamma}(\delta_{x,y}-\mathcal{G}^{+}(x;y+e_{2})-\mathcal{G}^{+}(x,y+e_{2}-e_{1}))f(y). (12)
  • Proof.  To prove the uniqueness result, it is sufficient to show that the corresponding homogeneous problem has only the trivial solution. Let uu be a radiating solution to the homogeneous problem 𝒫D\mathcal{P}_{D}. Then Theorem 4.2 immediately implies u(x)=0u(x)=0 for all xΩ̊x\in\mathring{\Omega}.

    To show the existence results, let us first check that u(x)u(x) satisfies the boundary condition (2b). Since 𝒢+(x;y)=0\mathcal{G}^{+}(x;y)=0 for all xΓx\in\Gamma and any yy we get

    u(x1,0)=y1δx1,y1δ0,0f(y)=f(x1).u(x_{1},0)=\sum_{y_{1}\in\mathbb{R}}\delta_{x_{1},y_{1}}\delta_{0,0}f(y)=f(x_{1}).

    Now let us check that u(x)u(x) satisfies the Helmholtz equation (2a). For x2>1x_{2}>1 all terms (Δd+k2)δx,y(\Delta_{d}+k^{2})\delta_{x,y}, (Δd+k2)𝒢+(x;y+e2)=δx,y+e2(\Delta_{d}+k^{2})\mathcal{G}^{+}(x;y+e_{2})=\delta_{x,y+e_{2}} and (Δd+k2)𝒢+(x;y+e2e1)=δx,y+e2e1(\Delta_{d}+k^{2})\mathcal{G}^{+}(x;y+e_{2}-e_{1})=\delta_{x,y+e_{2}-e_{1}} are equal to zero. Consequently, (Δd+k2)u(x)=0(\Delta_{d}+k^{2})u(x)=0 for points x=(x1,x2)x=(x_{1},x_{2}) with x2>1x_{2}>1. It remains to consider the case x2=1x_{2}=1. By the direct calculation we have

    (Δd+k2)δx,yf(y)\displaystyle(\Delta_{d}+k^{2})\delta_{x,y}f(y) =\displaystyle= f(x1)+f(x1+1),\displaystyle f(x_{1})+f(x_{1}+1),
    (Δd+k2)𝒢+(x;y+e2)f(y)\displaystyle(\Delta_{d}+k^{2})\mathcal{G}^{+}(x;y+e_{2})f(y) =\displaystyle= f(x1),\displaystyle f(x_{1}),
    (Δd+k2)𝒢+(x;y+e2e1)f(y)\displaystyle(\Delta_{d}+k^{2})\mathcal{G}^{+}(x;y+e_{2}-e_{1})f(y) =\displaystyle= f(x1+1),\displaystyle f(x_{1}+1),

    and as a result (Δd+k2)u(x)=0(\Delta_{d}+k^{2})u(x)=0. Thus, u(x)u(x) is a solution to (2a). Since 𝒢+\mathcal{G}^{+} satisfies the radiation condition and ff is supported on the finite subset of Γ\Gamma it follows that u(x)u(x) is the unique radiating solution to the problem (2a), (2b). \Box

5 Numerical results

The main issue for numerical evaluation of the solution (12) is to compute the lattice Green’s function. For this purpose we apply the method developed in [4]. Using 8-fold symmetry, we need only to compute the lattice Green’s function 𝒢(i,j)\mathcal{G}(i,j) with ij0i\geq j\geq 0. Following to [4], let us introduce the vectors 𝒱2p=(𝒢(2p,0),𝒢(2p1,1),,𝒢(p,p))\mathcal{V}_{2p}=(\mathcal{G}(2p,0),\mathcal{G}(2p-1,1),\dots,\mathcal{G}(p,p))^{\top} and 𝒱2p+1=(𝒢(2p+1,0),𝒢(2p,1),,𝒢(p+1,p))\mathcal{V}_{2p+1}=(\mathcal{G}(2p+1,0),\mathcal{G}(2p,1),\dots,\mathcal{G}(p+1,p))^{\top} that collect all distinct Green’s functions 𝒢(i,j)\mathcal{G}(i,j) with “Manhattan distances” i+j\mid i\mid+\mid j\mid of 2p2p and 2p+12p+1, respectively. For any Manhattan distance larger than 1, equation

(Δd+k2)𝒢(x)=δx,0(\Delta_{d}+k^{2})\mathcal{G}(x)=\delta_{x,0}

can be written in the matrix form γn(k)𝒱n=αn(k)𝒱n1+βn(k)𝒱n+1\gamma_{n}(k)\mathcal{V}_{n}=\alpha_{n}(k)\mathcal{V}_{n-1}+\beta_{n}(k)\mathcal{V}_{n+1} where αn(k)\alpha_{n}(k), βn(k)\beta_{n}(k) and γn(k)\gamma_{n}(k) are sparse matrices (cf., Appendix A). Notice that only the dimensions of these matrices depend on nn. It is shown in [4] that, for any n1n\geq 1, we have

𝒱n=An(k)𝒱n1,\mathcal{V}_{n}=A_{n}(k)\mathcal{V}_{n-1},

where the matrices An(k)A_{n}(k) are defined by the following recurrence formula

An(k)=[γn(k)βn(k)An+1]1αn(k).A_{n}(k)=[\gamma_{n}(k)-\beta_{n}(k)A_{n+1}]^{-1}\alpha_{n}(k).

They can be computed starting from a sufficiently large NN with AN+1(k)=0A_{N+1}(k)=0. Here, it is worth mentioning, that for k=2k=2 we need to choose a better “initial guess” than AN+1(k)=0A_{N+1}(k)=0, since in this case detγn(k)=0\mathrm{det}\gamma_{n}(k)=0 and the matrix γn(k)βn(k)An+1\gamma_{n}(k)-\beta_{n}(k)A_{n+1} is not invertible.

Once An(k)A_{n}(k) are known, we have 𝒱n=An(k)A1(k)𝒱0\mathcal{V}_{n}=A_{n}(k)\dots A_{1}(k)\mathcal{V}_{0}, where 𝒱0=𝒢(0,0)\mathcal{V}_{0}=\mathcal{G}(0,0). In particular, 𝒱1=𝒢(1,0)=A1(k)𝒢(0,0)\mathcal{V}_{1}=\mathcal{G}(1,0)=A_{1}(k)\mathcal{G}(0,0) which, together with 6𝒢(1,0)(6k2)𝒢(0,0)=16\mathcal{G}(1,0)-(6-k^{2})\mathcal{G}(0,0)=1, gives 𝒢(0,0)=1/[6A1(k)6+k2]\mathcal{G}(0,0)=1/[6A_{1}(k)-6+k^{2}]. This completes the calculation of the Green’s function using elementary operations and no integrals. Notice also one more important advantage of this method. The An(k)A_{n}(k) matrices are calculated coming down from asymptotically large Manhattan distances. As they are propagated towards smaller Manhattan distances, it definitely gives us the physical solution.

Finally, we demonstrate our theoretical and numerical approaches for the diffraction problem on triangular lattice. For this purpose we take k=2k=\sqrt{2} and consider the wave U(x1,x2,t)=exp(ιx2π/3ιωt)U(x_{1},x_{2},t)=\exp(\iota x_{2}\pi/3-\iota\omega t) on the semi-infinite lattice ×\mathbb{Z}\times\mathbb{Z}^{-} which encounters an obstacle at x2=0x_{2}=0 with two small openings formed by four nodes {(11,0),(10,0),(10,0),(11,0)}\{(-11,0),(-10,0),(10,0),(11,0)\}. It is easy to check that exp(ιx2π/3)\exp(\iota x_{2}\pi/3) satisfies the discrete Helmholtz equation. Thus, our goal is to evaluate numerically a radiating solution of Problem 𝒫D\mathcal{P}_{\mathrm{D}} where, for the given data, we take f(±10)=f(±11)=1f(\pm 10)=f(\pm 11)=1 and f(m)=0f(m)=0, m+\{11,10,10,11}m\in\mathbb{Z}^{+}\backslash\{-11,-10,10,11\}.

Refer to caption
(a) The density plot of Re𝒢\mathrm{Re}\,\mathcal{G}.
Refer to caption
(b) The density plot of Reu\mathrm{Re}\,u.
Refer to caption
(c) The graph of Reu(,253)\mathrm{Re}\,u(\cdot,25\sqrt{3}).
Figure 2: Plots are represented in original coordinates of the triangular lattice 𝔗\mathfrak{T}. k=2k=\sqrt{2}.

The results of numerical evaluations are plotted in Figure 2 in original coordinates of 𝔗\mathfrak{T}, where (a) shows the density plots of the Green’s function e𝒢\mathrm{\Re e}\,\mathcal{G}, (b) shows the density plots of cos(ιx2π/3)\cos(\iota x_{2}\pi/3) on the lower half-plane and eu\mathrm{\Re e}\,u for the case k=2k=\sqrt{2} on the upper half-plane, and (c) presents the graph of Reu(,253)\mathrm{Re}\,u(\cdot,25\sqrt{3}). Some key features of the numerical solution can be immediately observed, namely, as expected, the symmetry of eu\mathrm{\Re e}\,u and contributions from the wavefront that create a variable intensity.

6 Discussion

In this paper, we have constructed the discrete scattering theory for the two-dimensional discrete Helmholtz equation with a real wave number k(0,3)\{22}k\in(0,3)\backslash\{2\sqrt{2}\} for the semi-infinite triangular lattices. The main objective was to prove the unique solvability result and derive an exact formula for the solution to a Dirichlet problem. For simplicity we restricted ourselves to compact boundary data. For non-compact boundary data we should require some extra conditions at infinity, cf. [15].

Similarly to the continuum theory, we used the notion of the radiating solution for the continuous Helmholtz equation

Δu(x)+κ2u(x)=0,κ\{0},x2.\Delta u(x)+\kappa^{2}u(x)=0,\quad\kappa\in\mathbb{R}\backslash\{0\},\ x\in\mathbb{R}^{2}. (13)

Recall that a solution uu to the equation (13) is called radiating if it satisfies the Sommerfeld radiation condition

xu(x)ικu(x)=o(x12)\frac{\partial}{\partial\mid x\mid}u(x)-\iota\mid\kappa\mid u(x)=o(\mid x\mid^{-\frac{1}{2}}) (14)

for x\mid x\mid\to\infty uniformly in all directions x/xx/\mid x\mid. There are evident similarities and differences between conditions (8), (9) and (14). For example, we see the same decay at infinity, but, in contrast to (14), the second condition of (8) or (9) is anisotropic: the factor eιξj(α,k)e^{\iota\xi^{*}_{j}(\alpha,k)} is not a constant. Moreover, it is well known that the radiating solutions uu to the continuous Helmholtz equation automatically satisfy the Sommerfeld’s finiteness condition, however this fact is not shown for the discrete case so far. Therefore, the first condition of (8) or (9) is included in our definition.

In the present paper, the problems under consideration have an infinite boundary. Within the continuum framework it is well-known that, in general, when the surface is unbounded, we cannot neglect the contribution of that surface waves at infinity. In this case, the Sommerfeld radiation condition is no longer appropriate and a proper modification is needed. To the best of the authors’ knowledge, different radiation conditions are provided only for the half-plane (and locally perturbed half-plane) problems, cf. [7, 8, 9, 10, 11, 12], and finding radiation conditions for arbitrary wedge-shaped regions remains open. In case of the square lattice we have proposed sufficient conditions for the given boundary data at infinity, which ensures to have an unique radiation solution to the corresponding problem, cf. [15]. Here, in order to avoid further technical difficulties, we restricted ourselves with compactly supported data on the boundary.

Finally, let us note that comparing the results of problems in the continuous framework and results obtained in the continuum limit of discrete problems deserves the high interest of scientists, however, it is beyond the scope of this paper.

Acknowledgments

This work was supported by Shota Rustaveli National Science Foundation of Georgia (SRNSFG) [FR-21-301]

Appendix A Sparse matrices

The sparse matrices αn(k)\alpha_{n}(k), βn(k)\beta_{n}(k) and γn(k)\gamma_{n}(k) are defined as follows: if n=2pn=2p then α2p(k)\alpha_{2p}(k) is a (p+1)×p(p+1)\times p matrix such that α2p(k)i,i=1\alpha_{2p}(k)\mid_{i,i}=1, i=1,p¯i=\overline{1,p}, α2p(k)i,i1=1\alpha_{2p}(k)\mid_{i,i-1}=1, i=2,p¯i=\overline{2,p}, while α2p(k)p+1,p=2\alpha_{2p}(k)\mid_{p+1,p}=2, and all other matrix elements are zero. The β2p(k)\beta_{2p}(k) is a (p+1)×(p+1)(p+1)\times(p+1) matrix such that β2p(k)i,i=1\beta_{2p}(k)\mid_{i,i}=1, i=1,p¯i=\overline{1,p}, β2p(k)i,i+1=1\beta_{2p}(k)\mid_{i,i+1}=1, i=2,p¯i=\overline{2,p}, while β2p(k)p+1,p+1=β2p(k)1,2=2\beta_{2p}(k)\mid_{p+1,p+1}=\beta_{2p}(k)\mid_{1,2}=2, and all other matrix elements are zero. The γ2p(k)\gamma_{2p}(k) is a (p+1)×(p+1)(p+1)\times(p+1) matrix such that γ2p(k)i,i=6k2\gamma_{2p}(k)\mid_{i,i}=6-k^{2}, i=1,p+1¯i=\overline{1,p+1}, γ2p(k)i,i+1=γ2p(k)i,i1=1\gamma_{2p}(k)\mid_{i,i+1}=\gamma_{2p}(k)\mid_{i,i-1}=-1, i=2,p¯i=\overline{2,p}, and γ2p(k)1,2=γ2p(k)p+1,p=2\gamma_{2p}(k)\mid_{1,2}=\gamma_{2p}(k)\mid_{p+1,p}=-2.

If n=2p+1n=2p+1 then α2p+1(k)\alpha_{2p+1}(k) is a (p+1)×(p+1)(p+1)\times(p+1) matrix such that α2p+1(k)i,i=1\alpha_{2p+1}(k)\mid_{i,i}=1, i=1,p+1¯i=\overline{1,p+1}, α2p+1(k)i,i1=1\alpha_{2p+1}(k)\mid_{i,i-1}=1, i=2,p+1¯i=\overline{2,p+1}, and all other matrix elements are zero. The β2p+1(k)\beta_{2p+1}(k) is a (p+1)×(p+2)(p+1)\times(p+2) matrix such that β2p+1(k)i,i=1\beta_{2p+1}(k)\mid_{i,i}=1, i=1,p+1¯i=\overline{1,p+1}, β2p+1(k)i,i+1=1\beta_{2p+1}(k)\mid_{i,i+1}=1, i=2,p+1¯i=\overline{2,p+1}, while β2p+1(k)1,2=2\beta_{2p+1}(k)\mid_{1,2}=2, and all other matrix elements are zero. The γ2p(k)\gamma_{2p}(k) is a (p+1)×(p+1)(p+1)\times(p+1) matrix such that γ2p+1(k)i,i=6k2\gamma_{2p+1}(k)\mid_{i,i}=6-k^{2}, i=1,p¯i=\overline{1,p}, γ2p+1(k)p+1,p+1=5k2\gamma_{2p+1}(k)\mid_{p+1,p+1}=5-k^{2}, while γ2p+1(k)i,i+1=1\gamma_{2p+1}(k)\mid_{i,i+1}=-1, i=2,p¯i=\overline{2,p}, γ2p+1(k)i,i1=1\gamma_{2p+1}(k)\mid_{i,i-1}=-1, i=2,p+1¯i=\overline{2,p+1}, and γ2p+1(k)1,2=γ2p+1(k)p+1,p=2\gamma_{2p+1}(k)\mid_{1,2}=\gamma_{2p+1}(k)\mid_{p+1,p}=-2. Finally, γ1(k)\gamma_{1}(k) is a 1×11\times 1 matrixs with an element 4k24-k^{2}.

References

  • [1] Bhat H.S, Osting B. Diffraction on the Two-Dimensional Square Lattice, SIAM J. Appl. Math. 70 (2009), 1389-1406. https://doi.org/10.1137/080735345.
  • [2] Born M, Huang K. Dynamical Theory of Crystal Lattices. Clarendon Press, Oxford; 1954.
  • [3] Burke J.G. Origins of the Science of Crystals. University of California Press, Berklay and Los Angeles; 1966.
  • [4] Berciu M, Cook A.M. Efficient computation of lattice Green’s functions for models with nearest-neighbour hopping, Europhys. Lett. 2010; 92: 40003. DOI: 10.1209/0295-5075/92/40003.
  • [5] Brillouin L. Wave Propagation in Periodic Structures. Electric Filters and Crystal Lattices. International Series in Pure and Applied Physics: McGraw-Hill; 1946.
  • [6] Caloz C, Itoh T. Electromagnetic Metamaterials: Transmission Line theory and microwave applications: the engineering approach. John Wiley & Sons, Inc. Hoboken, New Jersey; 2006.
  • [7] Chandler-Wilde S.N., Boundary value problems for the Helmholtz equation in a half-plane, in: G. Cohen, (ed.), Proc. 3rd Int. Conf. on Mathematical and Numerical Aspects of Wave Propagation, SIAM, Philadelphia, 1995; 188-197.
  • [8] Chandler-Wilde S.N., The impedance boundary value problem for the Helmholtz equation in a half-plane, Math. Methods Appl. Sci. 20 (1997) 813-840. https://doi.org/10.1002/(SICI)1099-1476(19970710)20:10<<813::AID-MMA883>>3.0.CO;2-R.
  • [9] Chandler-Wilde S.N., Zhang B., A uniqueness result for scattering by infinite rough surfaces, SIAM J. Appl. Math. 58 (1998) 1774-1790. https://doi.org/10.1137/S0036139996309722.
  • [10] Chandler-Wilde S.N., Zhang B., Electromagnetic scattering by an inhomogeneous conducting or dielectric layer on a perfectly conducting plate, Proc. R. Soc. Lond. A. 454 (1998) 519-542. https://doi.org/10.1098/rspa.1998.0173.
  • [11] Durán M. , I. Muga, Nédélec J.C., The Helmholtz equation in a locally perturbed half-plane with passive boundary, IMA J. Appl. Math. 71 (2006) 853-876. https://doi.org/10.1093/imamat/hxl023.
  • [12] Durán M. , I. Muga, Nédélec J.C., The Helmholtz equation in a locally perturbed half-space with non-absorbing boundary, Arch. Ration. Mech. Anal. 191(2009) 143-172. https://doi.org/10.1007/s00205-008-0135-3.
  • [13] Dove MT. Structure and Dynamics: An Atomic View of Materials. Oxford University Press; 2002.
  • [14] Kapanadze D. The far-field behaviour of Green’s function for a triangular lattice and radiation conditions, Math Meth Appl Sci. 2021; 44(17):12746-12759. DOI: 10.1002/mma.7575
  • [15] Kapanadze D, Pesetskaya E. Diffraction problems for two-dimensional lattice waves in a quadrant, Wave Motion 2021; 100:102671. DOI: 10.1016/j.wavemoti.2020.102671
  • [16] Horiguchi T. Lattice Green’s Functions for the Triangular and Honeycomb Lattices, J. Math. Phys. 1972; 13(9):1411-1419. DOI: 10.1063/1.1666155
  • [17] Slepyan LI. Models and Phenomena in Fracture Mechanics. Springer, New York; 2002.